Next Article in Journal
The Cyclic Nitronate Route to Pharmaceutical Molecules: Synthesis of GSK’s Potent PDE4 Inhibitor as a Case Study
Previous Article in Journal
Development and Validation of a TLC-Densitometry Method for Histamine Monitoring in Fish and Fishery Products
Previous Article in Special Issue
Computational Prediction of Chiral Iron Complexes for Asymmetric Transfer Hydrogenation of Pyruvic Acid to Lactic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A DFT Study on FeI/FeII/FeIII Mechanism of the Cross-Coupling between Haloalkane and Aryl Grignard Reagent Catalyzed by Iron-SciOPP Complexes

by
Akhilesh K. Sharma
1 and
Masaharu Nakamura
1,2,*
1
International Research Center for Elements Science (IRCELS), Institute for Chemical Research, Kyoto University, Uji, Kyoto 611-0011, Japan
2
Department of Energy and Hydrocarbon Chemistry, Graduate School of Engineering, Kyoto University, Kyoto 615-8510, Japan
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(16), 3612; https://doi.org/10.3390/molecules25163612
Submission received: 1 July 2020 / Revised: 3 August 2020 / Accepted: 5 August 2020 / Published: 8 August 2020
(This article belongs to the Special Issue Recent Advances in Iron Catalysis)

Abstract

:
To explore plausible reaction pathways of the cross-coupling reaction between a haloalkane and an aryl metal reagent catalyzed by an iron–phosphine complex, we examine the reaction of FeBrPh(SciOPP) 1 and bromocycloheptane employing density functional theory (DFT) calculations. Besides the cross-coupling, we also examined the competitive pathways of β-hydrogen elimination to give the corresponding alkene byproduct. The DFT study on the reaction pathways explains the cross-coupling selectivity over the elimination in terms of FeI/FeII/FeIII mechanism which involves the generation of alkyl radical intermediates and their propagation in a chain reaction manner. The present study gives insight into the detailed molecular mechanic of the cross-coupling reaction and revises the FeII/FeII mechanisms previously proposed by us and others.

1. Introduction

Transition-metal-catalyzed cross-coupling reaction is one of the most versatile synthetic tools for constructing carbon frameworks of various functional molecules, e.g., pharmaceuticals, agrochemicals, and organic electronic materials. Iron catalysts have emerged as a sustainable alternative for conventional palladium and nickel catalysts and have attracted significant attention due to their practical merits of cost-effectiveness, low toxicity, and the abundance on the earth [1,2,3,4]. Despite the remarkable progress of iron-catalyzed cross-coupling reactions in organic synthesis [5,6,7,8,9,10,11,12,13,14], their reaction mechanisms remain elusive and attract considerable interests to their underlying molecular mechanisms [15,16,17,18,19].
More specifically, iron catalysts have proven useful for cross-coupling reactions of non-activated haloalkanes, a synthetically valuable yet challenging substrate, which is prone to undergo β-hydrogen elimination in the conventional palladium-catalyzed cross-couplings to form undesired alkene byproducts [20,21,22,23]. Numerous synthetic and mechanistic studies have thus appeared on this class of iron-catalyzed cross-coupling. The use of the controlled-addition technique of Grignard reagent in Kumada–Tamao-Corriu (KTC) coupling or using less reactive boron reagents in Suzuki–Miyaura (SM) coupling have proven, in combination with the use of bulky bisphosphine ligands, to be the successful strategy for the iron-catalyzed haloalkane coupling reactions [24,25,26,27].
There is a consensus that the haloalkane substrates undergo halogen abstraction by an (organo)iron species to generate the corresponding alkyl radical intermediates. A variety of radical mechanisms have been discussed thus far by Hu [28], Norrby [29], Bedford [30], Tonzetich [31], Fürstner [22], Neidig [32,33], and us [19,34]. Our group and the Neidig group have independently crystalized iron-bisphosphine complexes, such as FeIIX2(SciOPP), FeIIXAr(SciOPP), FeIIAr2(SciOPP), where X = halide and Ar = Ph or Mes [32,33,34]. Furthermore, we identified the presence of FeIIBrMes(SciOPP), FeIIMes2(SciOPP) in the reaction mixture of the cross-coupling reaction by using solution-phase synchrotron X-ray absorption spectroscopic (XAS) studies [34]. According to this study, we proposed an FeII/FeII mechanism based on identification of FeIIAr2(SciOPP) (Figure 1a, Ar = Mes and X = Br). The Neidig group also identified the presence of iron(II) complexes in the solution phase through intensive mechanistic studies, consisting of Mössbauer and magnetic circular dichroism (MCD) spectroscopy, and synthetic approaches [32,33]. They identified that FeIIXAr(SciOPP) is the active iron species of the cross-coupling reaction suppressing undesired β-hydrogen elimination reactions (Figure 1a, Ar = Ph). On the other hand, the Bedford group crystalized iron(I) species, [FeIBr(dpbz)2] which is formed from iron(II)-precatalyst and they proposed a low-coordinate iron(I)-phosphine species is involved in the cross-coupling reactions [35]. Identification of such low-coordinate iron(I)-complexes in the reaction mixture was not fully successful, and their involvement in the reaction mechanism remains elusive. The active iron-species that generate alkyl radical intermediates and the iron-species that react with the radical intermediate to effect the carbon–carbon bond formation have also remained unknown.
Recently, our group developed enantioselective cross-coupling reactions using chiral bisphosphine ligands and proposed that the radical species may not react with the iron(III)-species which takes part in the generation of the radical (in-cage mechanism) [36,37]. We further performed the detailed DFT analysis of the reaction of α-chloroesters with the Grignard reagent in presence of chiral bisphosphine ligand and found that FeI/FeII/FeIII mechanism operates in the reaction (Figure 1b) [38]. The iron(I) species activates the carbon–halogen bond of halo-ester substrates leading to the generation of α-ester-radical. In the next step, another molecule of FeIIXAr (active species) traps the generated radical (out-of-cage or chain mechanism) and lead to product formation through iron(III) species. Parallelly, the Gutierrez group reported DFT mechanistic study of the same reaction leading to the virtually same conclusion [39]. Despite these recent mechanistic insights, the use of electronically and sterically different iron-bisphosphine complexes, different substrates, and reagents in cross-coupling reactions make it difficult to generalize the mechanism understanding.
The iron-SciOPP catalysts have been known as a superior catalyst for the cross-coupling of alkyl halides and Grignard reagent [26]. Due to bulky nature and strong chelation ability of the ligand, it avoids the formation of multinuclear or multiligand iron–phosphine complexes and also suppresses the formation of undesired ate complexes, which presumably responsible for increased reactivity and selectivity by minimizing side reactions. The mechanistic studies by Neidig showed that FeIIBrPh(SciOPP) complex reacts with a haloalkane to give the corresponding cross-coupling product selectively, but the detailed mechanism has remained unclear [33]. In our pursuit to identify the iron species involved in the elementary steps of the iron-catalyzed haloalkane coupling, DFT is used to study the reaction between FeIIBrPh(SciOPP) and bromocycloheptane (Figure 1c), which has been believed as the key for the selective cross-coupling of a haloalkane with an aryl metal reagent in the presence of an iron-bisphosphine catalyst. The plausible reaction pathways, FeII/FeII, and FeI/FeII/FeIII are examined and compared in this study.

2. Results and Discussion

2.1. Reaction of FeIIBrPh(SciOPP) with Bromocycloheptane

2.1.1. Cycloheptyl Radical Generation via C–Br Activation by Iron(II)-Species

The reaction mechanism from FeIIBrPh(SciOPP) is studied as it is predicted to be the starting species in cross-coupling reactions. The TSs with lowest energy spin states as found in our previous study were only considered, due to large system size [38]. The first step is the C–Br activation of bromocycloheptane by quintet FeIIBrPh(SciOPP) (52PhBr), through an atom transfer mechanism (Figure 2). To a tetrahedral geometry of FeIIBrPh(SciOPP), 52PhBr, bromine-atom of substrate (S) approaches in the plane the same as P–Fe–P to form pre-reaction complex 52PhBr-RBr, which is 6.7 kcal/mol higher in energy than the separated molecules. The 5TS1 has a distorted square-pyramidal geometry, with the incoming bromine-atom positions in the square plane and the other bromine in the pyramidal position. The activation barrier of this process is 24.1 kcal/mol. The Mulliken spin densities on the iron center of the catalyst, and C1 and bromine of substrate are ρ(Fe) = 3.316, ρ (Br) = −0.055, and ρ(C1) = 0.672, respectively (Figure S1), indicating that during this step, quintet iron(II) is converted to quartet-iron(III) species with the generation of doublet cycloheptyl radical, 2. The formed 53PhBrBrR and 43PhBrBr species are respectively 21.7 and 14.9 kcal/mol higher in energy than 52PhBr, suggesting the reversible nature of the C–Br activation and will lead to a low concentration of radical 2.

2.1.2. Reaction of Cycloheptyl Radical with Iron(II)-Species: C–C Bond Formation and Generation of Iron(I) Species

The generated radical intermediate can react with either 43PhBrBr (iron(III)-species, in-cage mechanism) or 52PhBr (iron(II)-species, out-of-cage mechanism) present in the reaction mixture [36,38]. For the reaction of cycloheptyl radical () with iron(II) species, the first step involves the out-of-cage movement of 2 from the intermediate 43PhBrBrR. Then the 2 coordinates to the iron center through 4TS2 with a lower barrier than 5TS1 (20.7 kcal/mol from starting 52PhBr and 5.8 kcal/mol from the second 52PhBr, Figure 3). It leads to the formation of lower energy iron(III) species 43PhBrR, which is 11.7 kcal/mol lower in energy. The consequent reductive elimination forms C–C coupling product P1, which proceeds with the activation barrier of 13.1 kcal/mol (4TS3). The reductive elimination process is highly exergonic and the corresponding iron(I)-product complex (41Br-PhR) is 16.6 kcal/mol lower in energy than 43PhBrR. Afterwards, the product will decoordinate from 41Br-PhR leading to the formation of 5.7 kcal/mol higher energy species 41Br, probably through a barrierless process. The transient species 41Br readily reacts with a THF molecule (solvent) leading to the formation of lower energy species 41Br-THF (0.9 kcal/mol lower energy than 41Br-PhR). The whole process for the reaction of FeIIBrPh(SciOPP) with bromocycloheptane leading to the formation of product and 41Br-THF is exergonic by 13.3 kcal/mol.

2.1.3. Reaction of Cycloheptyl Radical with FeIIIBr2Ph(SciOPP) (In-Cage Mechanism)

The reaction of cycloheptyl radical with 43PhBrBr leading to a cross-coupling product turns out to have a higher free energy barrier of 26.9 kcal/mol (Figure S2). This process occurs through an outer-sphere fashion (5TS4), where attacks directly the ipso-carbon of the phenyl group and does not coordinate with the iron center. The reaction leading to byproduct alkene formation occurs by the simultaneous transfer of bromine and hydrogen from bromocycloheptane to iron and the ipso-carbon of the phenyl group of 52PhBr, respectively, via 5TS5 with a 25.2 kcal/mol energy barrier. Both processes for the formation of the cross-coupling and the alkene products are highly exothermic (ca. 35–45 kcal/mol). The formation of alkene is favorable kinetically over the cross-coupling. However, it does not agree with experimental observation.
Although the reaction of cycloheptyl radical with FeIIIBr2Ph(SciOPP), 43PhBrBr leading to cross-coupling product and alkene formation is highly exothermic, it is unlikely as 5TS4 and 5TS5 are 6.2 and 4.5 kcal/mol higher in energy than 4TS2. Hence, the cross-coupling reaction proceeds by the reaction of alkyl radical with iron(II)-species. Similar to our previous study, iron-SciOPP complexes also do not favor FeII/FeIII mechanism.

2.1.4. Reaction of Cycloheptyl Radical with Iron(II)-Species: Alkene Byproduct Formation

From complex 43PhBrR the possibility of formation of alkene is also considered (Figure 4). It involves the transfer of β-hydrogen from a coordinated cycloheptyl group to phenyl ligand (4TS6). The activation barrier for this process is apparently high and 4TS6 is 14.1 kcal/mol higher in energy than 4TS3 and the formation of alkene is unlikely from 43PhBrR. The TSs for coupling and alkene formation by the outer-sphere reaction of radical with phenyl of iron(II) species, 52PhBr could not be located.
Further, we checked the possibility of alkene formation by H-atom transfer (4TS7) from cycloheptyl radical to iron-atom of the iron(II) complex, 52PhBr. The activation barrier for this process is 12.9 kcal/mol. It leads to the formation of the iron(III)-hydride complex, 43PhBrH. The following reductive elimination between the hydride and phenyl groups through 4TS8 lead to the formation of benzene. This pathway is unfavorable as 4TS7 is 7.7 kcal/mol higher in energy than 4TS2. Hence, the reaction of the alkyl radical with FeIIBrPh(SciOPP) does not give the alkene byproduct, which is in concert with the experimental observations.

2.1.5. Reaction of Iron(I) Species with Bromocycloheptane: Regeneration of FeIIBr2(SciOPP) and Cycloheptyl Radical

The iron(I) species formed during the reaction can react either with another iron(III) species to undergo a comproportionation to generate two iron(II) species, or with haloalkane substrate (of which concentration is larger than iron(III) species) to generate one iron(II) species and an alkyl radical intermediate. For the reaction to proceed the first step is the removal of THF from 41Br-THF to form 41Br and this process is endergonic by 6.6 kcal/mol (Figure 5). This step is required for both comproportionation and reaction with the haloalkane substrate. The coordination of the substrate through bromide to 41Br, leading to the formation of 41Br-BrR is exergonic by 2.3 kcal/mol. The activation barrier for the Br-atom transfer (4TS9) from 41Br-BrR species is 6.8 kcal/mol (barrier from 41Br-THF). The process is highly exergonic and leads to formation of cycloheptyl radical and iron(II) species 52BrBr. The generated 52BrBr species will undergo transmetalation and the radical species will react with the available 52PhBr in the reaction mixture, propagating the radical chain reaction [40]. Hence, the reaction will proceed through the out-of-cage mechanism, and similar to our previous study, the current reaction also follows the FeI/FeII/FeIII pathway [38].
The comproportionation between 41Br and 43PhBrBr (generated by substrate reaction with iron(II) species 52PhBr) may be a barrierless process and is a highly exergonic process (37.2 kcal/mol). Since the concentration of the haloalkane substrate is high enough under the catalytic coupling condition, the reaction between the iron(I) species and haloalkane is more likely. Hence, iron(I) species is the active species after its generation during the reaction, which reduces the activation barrier for C–Cl activation from 24.1 to 6.8 kcal/mol.

2.2. Proposed Catalytic Cycle Based on FeI/FeII/FeIII Mechanism

Based on the above theoretical studies, we propose the reaction mechanism depicted in Figure 6. The FeIIXAr(SciOPP) (52PhBr) is first converted to FeIXTHF(SciOPP) (41Br-THF) with the formation of the cross-coupling product. Afterwards, iron(I) species (41Br-THF) mainly activate the alky halide. Transmetalation of 41Br-THF is less likely due to the low concentration of Grignard reagent, which excludes the possibility of involvement of FeIArTHF(SciOPP) species. Hence, the catalytic cycle starts with the halogen atom abstraction from the alkyl halide (Alkyl-X) by iron(I) species (41Br) affording a dihaloiron species FeIIX2(SciOPP) (52BrBr) along with an alkyl radical intermediate (2) (Figure 6 and Figure S3). The resultant alkyl radical 2 escapes from the solvent cage and is further trapped by iron(II) species, FeIIXAr(SciOPP) (52PhBr) (which have a higher concentration) via the out-of-cage pathway. It led to the formation of iron(III) species FeIIIXArR(SciOPP) (43PhBrR). The reductive elimination from 43PhBrR led to the formation of a cross-coupling product and regenerating iron(I) species (41Br-THF). Based on this mechanism, the detection of both iron(II) and iron(I) species, and their roles in the cross-coupling reaction can be reasonably explained.

3. Computational Methods

The Gaussian09 program was used for quantum mechanical calculations reported in the study [41]. The DFT method B3LYP [42,43] was used for geometry optimization and energy calculations with the inclusion of solvent effects by the PCM method [44]. Grimme’s D3 dispersion correction method was used to include dispersion correction [45]. For geometry optimization, the 6-31G* basis set [46] was used for all atoms as it gave good agreement with the crystal structure of FeIIBrPh(SciOPP) and Fe IIBr2(SciOPP) (Figure S4). The geometry optimization with SDD [47] basis set for Fe, Br and 6-31G* for other atoms lead to relatively longer metal ligand distances. Hence, the 6-31G* basis set was used. The hessian calculation was performed to confirm minima (no imaginary frequency) and TS (one imaginary frequency). All TS were confirmed by visual inspection of imaginary frequency. The connectivity between TSs and minima were confirmed by IRC [48] calculations for 20 steps and subsequent optimization of the geometry at PCM/B3LYP/6-31G* level of theory.
The energies were further computed using the 6-311G** basis set [49] for all atoms and with Douglas–Kroll–Hess second-order scalar relativistic effects. The EZPE reported in the paper is calculated by adding zero-point-energy obtained at the 6-31G* basis set to electronic energy of higher basis set (EZPE = Eelec(6-311G**) + ZPE(6-31G*)). The free energies were calculated by adding thermal free energy correction obtained at 6-31G* to electronic energy of the 6-311G** basis set (G = Eelec(6-311G**) + GCorr(6-31G*)). The free energies were calculated at 298.15 K temperature and 1 atm pressure. Concisely, the energies reported in the paper were calculated at PCMTHF/B3LYP-D3/6-311G**//PCMTHF/B3LYP-D3/6-31G* level of theory.
The spin-state energies of iron complexes computed by a DFT method depend on the choice of the exchange-correlation functional [50,51]. Various functionals were recommended in the literature, while no consensus has been made for the optimal choice of the DFT method. Hence, care should be taken in choosing functional, especially for iron complexes with the small spin-state energy gaps. The DFT method (B3LYP-D3 Functional) used in this work predicted quintet as the ground state for iron(II) complexes, FeIIBrPh(SciOPP) and FeIIBr2(SciOPP), which is in agreement with the experimentally known spin state [33,34]. Additionally, the spin state for related iron(I) complex (Fe(η2-[TIPS-CC-H])Br(SciOPP)) was also predicted correctly [52]. Further, the OBPE-D3 functional (which is known to provide reliable spin state gap) [53] gave consistent results with B3LYP-D3 for the ground spin state (quartet) of iron(I) and iron(III) complexes (Table S1). The OPBE-D3 functional failed to give reliable results for the spin state of the FeIIBrPh(SciOPP) complex. Hence, the B3LYP-method reliably predicts the spin state of the current system and is used in this study. Further, we have also checked the effect of long-range-corrected (LC) exchange functional on barrier height [54] using CAM-B3LYP-D3 functional. The CAM-B3LYP-D3 functional led to a similar conclusion as of B3LYP-D3, and some TSs showed a higher energy barrier (details are given in Table S2).

4. Conclusions

We have studied the mechanism of the reaction between FeIIBrPh(SciOPP) and bromocycloheptane by using DFT methods and identified iron species of different oxidation states involved in a catalytic cycle of the iron-bisphosphine-catalyzed haloalkane couplings. The reaction mechanism follows the FeI/FeII/FeIII pathway, where iron(I) activates the C–Br bond of the haloalkane substrate, another molecule of FeIIBrPh(SciOPP) traps the generated alkyl radical species to form iron(III) species leading to the formation of the cross-coupling product and iron(I) species. The iron(I) species then again react with haloalkane generating alkyl radical and the reaction propagates through a radical chain mechanism. The current mechanism mirrors our previous mechanistic study on a chiral bisphosphine-catalyzed enantioselective cross-coupling despite the change of bisphosphine ligand from BenzP* to SciOPP.
The reaction of haloalkane with iron(II) species and the reaction of the alkyl radical with iron(III) species have a higher energy barrier than those of Iron(I) species and iron(II) species, respectively. The obtained energetics exclude the possibility of FeII/FeII and FeII/FeIII pathways proposed previously. Furthermore, we have found that an alkyl radical prefers to coordinate to the FeIIBrPh(SciOPP) species, instead of the β-hydrogen-atom abstraction ending with the alkene byproduct. The generated FeIIIBrPh(alkyl)(SciOPP) species favors reductive elimination to give the corresponding coupling product over the alkene byproduct. These results deepen our understanding of the iron-SciOPP catalysis and provide insights into the probable root for the side product formations. Further understanding of the mechanism of alkene formation with FeIIPh2(SciOPP) catalyst will help design better catalysts and synthetic methodology. The theoretical and experimental research in this line is now actively pursued in our laboratory and will be reported in due course.

Supplementary Materials

The following are available online. Cartesian coordinates of optimized geometries and their energies, Figure S1: Mulliken spin densities of the stationary points, Figure S2: Free energy profile for the reaction of FeII/FeIII mechanism, Figure S3: Energy profile for FeI/FeII/FeIII pathway starting the reaction from iron(I) species, Figure S4: Comparison of geometrical parameters of X-ray crystal structure of 2PhBr and 2BrBr with the optimized geometry in quintet spin state, Table S1: The comparison of energies for the iron(I), iron(II) and iron(III) complexes in different spin states at B3LYP-D3 and OPBE-D3 functionals. Table S2: Comparison of free energy barriers of catalytic cycle at B3LYP-D3 and CAM-B3LYP-D3 functionals.

Author Contributions

Conceptualization, M.N. and A.K.S.; methodology, A.K.S.; validation, A.K.S. and M.N.; writing—original draft preparation, A.K.S. and M.N.; writing—review and editing, M.N.; visualization, A.K.S.; supervision, M.N.; project administration, M.N.; funding acquisition, M.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a Grant-in-Aid for Scientific Research (20H02740) and Grant-in-Aid for Scientific Research on Innovative Areas “Synergistic Effects for Creation of Functional Molecules” (JP18064006) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT), and through the “Funding Program for Next generation World-Leading Researchers (Next Program)” initiated by the Council for Science and Technology Policy (CSTP). This work was supported by Integrated Research Consortium on Chemical Sciences. M.N. thanks for the financial support from Nissan Chemical Corporation and Tosoh Finechem Corporation.

Acknowledgments

Supercomputing facility at the Institute of Molecular Science in Japan and the Academic Center for Computing at Media Studies at Kyoto University in Japan are acknowledged for computational resources. Computation time was also provided by the Super Computer System, Institute for Chemical Research, Kyoto University. WMC Sameera and Miho Isegawa are also acknowledged for useful discussion and contributions in computational design. This work was also supported by the Integrated Research Consortium on Chemical Sciences (IRCCS).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Guðmundsson, A.; Bäckvall, J.-E. On the use of iron in organic chemistry. Molecules 2020, 25, 1349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Fürstner, A. Iron catalysis in organic synthesis: A critical assessment of what it takes to make this base metal a multitasking champion. ACS Cent. Sci. 2016, 2, 778–789. [Google Scholar] [CrossRef] [PubMed]
  3. Bauer, I.; Knölker, H.-J. Iron catalysis in organic synthesis. Chem. Rev. 2015, 115, 3170–3387. [Google Scholar] [CrossRef]
  4. Bolm, C.; Legros, J.; Le Paih, J.; Zani, L. Iron-catalyzed reactions in organic synthesis. Chem. Rev. 2004, 104, 6217–6254. [Google Scholar] [CrossRef] [PubMed]
  5. Piontek, A.; Bisz, E.; Szostak, M. Iron-catalyzed cross-couplings in the synthesis of pharmaceuticals: In pursuit of sustainability. Angew. Chem. Int. Ed. 2018, 57, 11116–11128. [Google Scholar] [CrossRef]
  6. Legros, J.; Figadere, B. Iron-promoted C–C bond formation in the total synthesis of natural products and drugs. Nat. Prod. Rep. 2015, 32, 1541–1555. [Google Scholar] [CrossRef] [Green Version]
  7. Cahiez, G.; Knochel, P.; Kuzmina, O.; Steib, A.; Moyeux, A. Recent advances in iron-catalyzed Csp2–Csp2 cross-couplings. Synthesis 2015, 47, 1696–1705. [Google Scholar] [CrossRef]
  8. Ilies, L.; Nakamura, E. Iron-catalyzed cross-coupling reactions. PATAI’S Chem. Funct. Groups 2012, 83, 1–209. [Google Scholar] [CrossRef]
  9. Guérinot, A.; Cossy, J. Iron-catalyzed C–C cross-couplings using organometallics. Top. Curr. Chem. 2016, 374, 49. [Google Scholar] [CrossRef]
  10. Tamura, M.; Kochi, J.K. Vinylation of grignard reagents. Catalysis by iron. J. Am. Chem. Soc. 1971, 93, 1487–1489. [Google Scholar] [CrossRef]
  11. Reddy, C.K.; Knochel, P. New cobalt and iron-catalyzed reactions of organozinc compounds. Angew. Chem. Int. Ed. 1996, 35, 1700–1701. [Google Scholar] [CrossRef]
  12. Cahiez, G.; Avedissian, H. Highly stereo and chemoselective iron-catalyzed alkenylation of organomagnesium compounds. Synthesis 1998, 1998, 1199–1205. [Google Scholar] [CrossRef]
  13. Fürstner, A.; Leitner, A.; Méndez, M.; Krause, H. Iron-catalyzed cross-coupling reactions. J. Am. Chem. Soc. 2002, 124, 13856–13863. [Google Scholar] [CrossRef] [PubMed]
  14. Hatakeyama, T.; Nakamura, M. Iron-catalyzed selective biaryl coupling: Remarkable suppression of homocoupling by the fluoride anion. J. Am. Chem. Soc. 2007, 129, 9844–9845. [Google Scholar] [CrossRef] [PubMed]
  15. Bedford, R.B. How low does iron go? Chasing the active species in Fe-catalyzed cross-coupling reactions. Acc. Chem. Res. 2015, 48, 1485–1493. [Google Scholar] [CrossRef] [PubMed]
  16. Neidig, M.L.; Carpenter, S.H.; Curran, D.J.; DeMuth, J.C.; Fleischauer, V.E.; Iannuzzi, T.E.; Neate, P.G.N.; Sears, J.D.; Wolford, N.J. Development and evolution of mechanistic understanding in iron-catalyzed cross-coupling. Acc. Chem. Res. 2018, 52, 140–150. [Google Scholar] [CrossRef]
  17. Parchomyk, T.; Koszinowski, K. Iron-catalyzed cross-coupling: Mechanistic insight for rational applications in synthesis. Synthesis 2017, 49, 3269–3280. [Google Scholar] [CrossRef]
  18. Mako, T.L.; Byers, J.A. Recent advances in iron-catalysed cross coupling reactions and their mechanistic underpinning. Inorg. Chem. Front. 2016, 3, 766–790. [Google Scholar] [CrossRef]
  19. Noda, D.; Sunada, Y.; Hatakeyama, T.; Nakamura, M.; Nagashima, H. Effect of TMEDA on iron-catalyzed coupling reactions of ArMgX with Alkyl halides. J. Am. Chem. Soc. 2009, 131, 6078–6079. [Google Scholar] [CrossRef]
  20. Nakamura, M.; Matsuo, K.; Ito, S.; Nakamura, E. Iron-catalyzed cross-coupling of primary and secondary Alkyl halides with Aryl grignard reagents. J. Am. Chem. Soc. 2004, 126, 3686–3687. [Google Scholar] [CrossRef]
  21. Nagano, T.; Hayashi, T. Iron-catalyzed grignard cross-coupling with Alkyl halides possessing β-hydrogens. Org. Lett. 2004, 6, 1297–1299. [Google Scholar] [CrossRef] [PubMed]
  22. Martin, R.; Fürstner, A. Cross-coupling of Alkyl halides with Aryl grignard reagents catalyzed by a low-valent iron complex. Angew. Chem. Int. Ed. 2004, 43, 3955–3957. [Google Scholar] [CrossRef] [PubMed]
  23. Bedford, R.B.; Bruce, D.W.; Frost, R.M.; Goodby, J.W.; Hird, M. Iron(III) salen-type catalysts for the cross-coupling of aryl grignards with alkyl halides bearing β-hydrogens. Chem. Commun. 2004, 2822. [Google Scholar] [CrossRef] [PubMed]
  24. Bedford, R.B.; Huwe, M.; Wilkinson, M.C. Iron-catalysed Negishi coupling of benzylhalides and phosphates. Chem. Commun. 2009, 600–602. [Google Scholar] [CrossRef]
  25. Hatakeyama, T.; Kondo, Y.; Fujiwara, Y.-I.; Takaya, H.; Ito, S.; Nakamura, E.; Nakamura, M. Iron-catalysed fluoroaromatic coupling reactions under catalytic modulation with 1,2-bis (diphenylphosphino) benzene. Chem. Commun. 2009, 1216–1218. [Google Scholar] [CrossRef]
  26. Hatakeyama, T.; Hashimoto, T.; Kondo, Y.; Fujiwara, Y.; Seike, H.; Takaya, H.; Tamada, Y.; Ono, T.; Nakamura, M. Iron-catalyzed Suzuki−Miyaura coupling of Alkyl halides. J. Am. Chem. Soc. 2010, 132, 10674–10676. [Google Scholar] [CrossRef]
  27. Hatakeyama, T.; Fujiwara, Y.-I.; Okada, Y.; Itoh, T.; Hashimoto, T.; Kawamura, S.; Ogata, K.; Takaya, H.; Nakamura, M. Kumada–Tamao–Corriu coupling of Alkyl halides catalyzed by an iron–bisphosphine complex. Chem. Lett. 2011, 40, 1030–1032. [Google Scholar] [CrossRef]
  28. Bauer, G.; Wodrich, M.D.; Scopelliti, R.; Hu, X. Iron pincer complexes as catalysts and intermediates in Alkyl–Aryl Kumada coupling reactions. Organometallics 2014, 34, 289–298. [Google Scholar] [CrossRef] [Green Version]
  29. Hedström, A.; Izakian, Z.; Vreto, I.; Wallentin, C.-J.; Norrby, P.-O. On the radical nature of iron-catalyzed cross-coupling reactions. Chem. A Eur. J. 2015, 21, 5946–5953. [Google Scholar] [CrossRef]
  30. Bedford, R.B.; Brenner, P.B.; Richards, E.; Carvell, T.W.; Cogswell, P.M.; Gallagher, T.; Harvey, J.; Murphy, D.; Neeve, E.C.; Nunn, J.; et al. Expedient iron-catalyzed coupling of Alkyl, Benzyl and Allyl halides with Arylboronic esters. Chem. A Eur. J. 2014, 20, 7935–7938. [Google Scholar] [CrossRef]
  31. Przyojski, J.A.; Veggeberg, K.P.; Arman, H.D.; Tonzetich, Z.J. Mechanistic studies of catalytic carbon–carbon cross-coupling by well-defined iron NHC complexes. ACS Catal. 2015, 5, 5938–5946. [Google Scholar] [CrossRef]
  32. Daifuku, S.L.; Al-Afyouni, M.H.; Snyder, B.E.R.; Kneebone, J.L.; Neidig, M.L. A combined Moössbauer, magnetic circular dichroism, and density functional theory approach for iron cross-coupling catalysis: Electronic structure, In Situ Formation, and reactivity of iron-Mesityl-Bisphosphines. J. Am. Chem. Soc. 2014, 136, 9132–9143. [Google Scholar] [CrossRef] [PubMed]
  33. Daifuku, S.L.; Kneebone, J.L.; Snyder, B.E.R.; Neidig, M.L. Iron (II) active species in iron–bisphosphine catalyzed Kumada and Suzuki–Miyaura cross-couplings of Phenyl Nucleophiles and secondary Alkyl halides. J. Am. Chem. Soc. 2015, 137, 11432–11444. [Google Scholar] [CrossRef] [PubMed]
  34. Takaya, H.; Nakajima, S.; Nakagawa, N.; Isozaki, K.; Iwamoto, T.; Imayoshi, R.; Gower, N.J.; Adak, L.; Hatakeyama, T.; Honma, T.; et al. Investigation of organoiron catalysis in Kumada–Tamao–Corriu-type cross-coupling reaction assisted by solution-phase X-ray absorption spectroscopy. Bull. Chem. Soc. Jpn. 2015, 88, 410–418. [Google Scholar] [CrossRef] [Green Version]
  35. Adams, C.J.; Bedford, R.B.; Carter, E.; Gower, N.J.; Haddow, M.; Harvey, J.; Huwe, M.; Cartes, M.Á.; Mansell, S.M.; Mendoza, C.; et al. Iron(I) in Negishi cross-coupling reactions. J. Am. Chem. Soc. 2012, 134, 10333–10336. [Google Scholar] [CrossRef]
  36. Jin, M.; Adak, L.; Nakamura, M. Iron-catalyzed enantioselective cross-coupling reactions of α-chloroesters with Aryl grignard reagents. J. Am. Chem. Soc. 2015, 137, 7128–7134. [Google Scholar] [CrossRef] [Green Version]
  37. Iwamoto, T.; Okuzono, C.; Adak, L.; Jin, M.; Nakamura, M. Iron-catalysed enantioselective Suzuki–Miyaura coupling of racemic alkyl bromides. Chem. Commun. 2019, 55, 1128–1131. [Google Scholar] [CrossRef]
  38. Sharma, A.K.; Sameera, W.M.C.; Jin, M.; Adak, L.; Okuzono, C.; Iwamoto, T.; Kato, M.; Nakamura, M.; Morokuma, K. DFT and AFIR study on the mechanism and the origin of enantioselectivity in iron-catalyzed cross-coupling reactions. J. Am. Chem. Soc. 2017, 139, 16117–16125. [Google Scholar] [CrossRef]
  39. Lee, W.; Zhou, J.; Gutierrez, O. Mechanism of Nakamura’s Bisphosphine-iron-catalyzed asymmetric C(sp2)–C(sp3) cross-coupling reaction: The role of spin in controlling Arylation pathways. J. Am. Chem. Soc. 2017, 139, 16126–16133. [Google Scholar] [CrossRef]
  40. Schley, N.D.; Fu, G.C. Nickel-catalyzed Negishi Arylations of Propargylic bromides: A mechanistic investigation. J. Am. Chem. Soc. 2014, 136, 16588–16593. [Google Scholar] [CrossRef] [Green Version]
  41. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.01/E.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  42. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef] [PubMed]
  45. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Ditchfield, R. Self-consistent molecular-orbital methods. IX. An extended Gaussian-type basis for molecular-orbital studies of organic molecules. J. Chem. Phys. 1971, 54, 724. [Google Scholar] [CrossRef]
  47. Dolg, M.; Wedig, U.; Stoll, H.; Preuss, H. Energy-adjusted ab initio pseudopotentials for the first row transition elements. J. Chem. Phys. 1987, 86, 866–872. [Google Scholar] [CrossRef]
  48. Fukui, K. The path of chemical reactions—The IRC approach. Acc. Chem. Res. 1981, 14, 363–368. [Google Scholar] [CrossRef]
  49. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  50. Radoń, M. Benchmarking quantum chemistry methods for spin-state energetics of iron complexes against quantitative experimental data. Phys. Chem. Chem. Phys. 2019, 21, 4854–4870. [Google Scholar] [CrossRef]
  51. Siig, O.S.; Kepp, K.P. Iron(II) and Iron(III) spin crossover: Toward an optimal density functional. J. Phys. Chem. A 2018, 122, 4208–4217. [Google Scholar] [CrossRef] [Green Version]
  52. Kneebone, J.L.; Brennessel, W.W.; Neidig, M.L. Intermediates and reactivity in iron-catalyzed cross-couplings of Alkynyl grignards with Alkyl Halides. J. Am. Chem. Soc. 2017, 139, 6988–7003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Swart, M. Accurate spin-state energies for iron complexes. J. Chem. Theory Comput. 2008, 4, 2057–2066. [Google Scholar] [CrossRef] [PubMed]
  54. Song, J.-W.; Hirosawa, T.; Tsuneda, T.; Hirao, K. Long-range corrected density functional calculations of chemical reactions: Redetermination of parameter. J. Chem. Phys. 2007, 126, 154105. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds are not available from the authors.
Figure 1. (a) Proposed catalytic cycle for cross-coupling between an alkyl halide and Grignard reagent catalyzed by iron-SciOPP complexes. (b) Proposed catalytic cycle based on a recent study involving iron-BenzP* catalyzed cross-coupling between α-chloroesters and Grignard reagent. (c) The iron-SciOPP-catalyzed cross-coupling reaction chosen for the current density functional theory (DFT) mechanistic study [26,27].
Figure 1. (a) Proposed catalytic cycle for cross-coupling between an alkyl halide and Grignard reagent catalyzed by iron-SciOPP complexes. (b) Proposed catalytic cycle based on a recent study involving iron-BenzP* catalyzed cross-coupling between α-chloroesters and Grignard reagent. (c) The iron-SciOPP-catalyzed cross-coupling reaction chosen for the current density functional theory (DFT) mechanistic study [26,27].
Molecules 25 03612 g001
Figure 2. Free energy profile for the reaction of iron(II) species with bromocycloheptane. The relative free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol with total electronic energies including ZPE (zero-point energy) correction (ΔEZPE) are given in parenthesis.
Figure 2. Free energy profile for the reaction of iron(II) species with bromocycloheptane. The relative free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol with total electronic energies including ZPE (zero-point energy) correction (ΔEZPE) are given in parenthesis.
Molecules 25 03612 g002
Figure 3. Free energy profile for cross-coupling by the reaction of iron(II) species with cycloheptyl radical, where radical is generated by the reaction of iron(II) species and bromocycloheptane. Free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Figure 3. Free energy profile for cross-coupling by the reaction of iron(II) species with cycloheptyl radical, where radical is generated by the reaction of iron(II) species and bromocycloheptane. Free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Molecules 25 03612 g003
Figure 4. Free energy profile for olefin formation by reaction of FeII species with cycloheptyl radical, where radical is generated by the reaction of iron(II) species and bromocycloheptane. Free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Figure 4. Free energy profile for olefin formation by reaction of FeII species with cycloheptyl radical, where radical is generated by the reaction of iron(II) species and bromocycloheptane. Free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Molecules 25 03612 g004
Figure 5. Free energy profile for probable reaction path of generated iron(I) species. Free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Figure 5. Free energy profile for probable reaction path of generated iron(I) species. Free energies with respect to 52PhBr and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Molecules 25 03612 g005
Figure 6. Revised catalytic cycle of iron-SciOPP catalyzed-KTC coupling of a haloalkane based on the current DFT study. Free energies with respect to 41Br-THF and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Figure 6. Revised catalytic cycle of iron-SciOPP catalyzed-KTC coupling of a haloalkane based on the current DFT study. Free energies with respect to 41Br-THF and bromocycloheptane are given in kcal/mol (ΔEZPE are given in parenthesis).
Molecules 25 03612 g006

Share and Cite

MDPI and ACS Style

Sharma, A.K.; Nakamura, M. A DFT Study on FeI/FeII/FeIII Mechanism of the Cross-Coupling between Haloalkane and Aryl Grignard Reagent Catalyzed by Iron-SciOPP Complexes. Molecules 2020, 25, 3612. https://doi.org/10.3390/molecules25163612

AMA Style

Sharma AK, Nakamura M. A DFT Study on FeI/FeII/FeIII Mechanism of the Cross-Coupling between Haloalkane and Aryl Grignard Reagent Catalyzed by Iron-SciOPP Complexes. Molecules. 2020; 25(16):3612. https://doi.org/10.3390/molecules25163612

Chicago/Turabian Style

Sharma, Akhilesh K., and Masaharu Nakamura. 2020. "A DFT Study on FeI/FeII/FeIII Mechanism of the Cross-Coupling between Haloalkane and Aryl Grignard Reagent Catalyzed by Iron-SciOPP Complexes" Molecules 25, no. 16: 3612. https://doi.org/10.3390/molecules25163612

Article Metrics

Back to TopTop